Вы находитесь на странице: 1из 9

Crack characterization using guided circumferential waves

Christine Valle
Department of Mechanical Engineering, University of Maine, Orono, Maine 04469-5711

Marc Niethammer
School of Civil and Environmental Engineering, Georgia Institute of Technology, Atlanta,
Georgia 30332-0355

Jianmin Qu
G. W. W. School of Mechanical Engineering, Georgia Institute of Technology, Atlanta, Georgia 30332-0405

Laurence J. Jacobs
School of Civil and Environmental Engineering, Georgia Institute of Technology, Atlanta,
Georgia 30332-0355

Received 1 August 2000; revised 15 March 2001; accepted 14 May 2001


This paper examines the propagation of guided circumferential waves in a hollow isotropic cylinder
that contains a crack, with the goal of using these guided waves to both locate and size the crack.
The crack is sized using a modified Aulds formula, which relates the cracks length to a reflected
energy coefficient. The crack is then located by operating on the backscattered signal with a
time-frequency digital signal processing DSP technique, and then comparing these results to those
obtained if the cylinder is perfect. The guided circumferential waves are generated with a
commercial finite element method FEM code. One objective of this work is to demonstrate the
effectiveness of using sophisticated DSP techniques to describe the effect of scattering on dispersive
waves, showing it is possible to characterize cracks systematically and accurately by quantifying
this scattering effect. The results show that the need for high frequency signals to detect small cracks
is significantly decreased by using these techniques. 2001 Acoustical Society of America.
DOI: 10.1121/1.1385899
PACS numbers: 43.20.Fn, 43.20.Mv DEC

I. INTRODUCTION

It is well known that fatigue cracks can initiate and grow


in the radial direction of an annular structure that is subjected
to a large number of fatigue cycles. These annular structures
are used extensively in a variety of industrial applications,
such as the aerospace, oil and nuclear industries. A quantitative and systematic inspection methodology is needed to detect and characterize these cracks, before a catastrophic
structural failure occurs.
Ultrasonic testing is a candidate nondestructive evaluation NDE methodology for this application. Unfortunately,
traditional ultrasonic techniques such as pulse-echo are ineffective in this application because of problems associated
with curvature and accessibility difficulties. Guided ultrasonic waves, i.e., waves that propagate in the direction of the
layer while behaving as standing waves through the thickness of the layer, are a potential alternative. Nagy et al.1
recently proposed using guided ultrasonic waves that propagate in the circumferential direction to detect radial cracks in
weep holes of airframes.
The main advantage in using guided ultrasonic waves is
that they can interrogate the entire specimen, including inaccessible portions. Unfortunately, the detected ultrasonic signals are very complicated, causing difficulties in signal interpretation. Previous researchers have studied guided waves in
plates2 and cylinders.3 For example, Alleyne and Cawley3
examine cylindrical waves that propagate down the axis of
the cylinder but remain standing in its circumferential direc1282

J. Acoust. Soc. Am. 110 (3), Pt. 1, Sep. 2001

tion to detect cracks in long thin pipes. That is, the expressions for the field quantities contain terms such as cos(n) or
sin(n), where is the usual polar angle see Fig. 1 and n is
an integer. Thus quantities such as velocities or stresses are
modulo 2 periodic. However, these longitudinal waves
are not well suited for determining the radial location of a
crack in a short annular structure. This is especially true for
a cylindrical component whose diameter is on the same order
of magnitude as its length. Circumferential waves, that is
guided waves that propagate in the circumferential direction
of the cylinder, i.e., in the direction see Fig. 1, are proposed for these applications. Such waves, contrary to the
previous case, contain terms such as e ik , where has the
same meaning as previously and k is a real number and not
an integer. Thus quantities such as velocities and stresses
are not periodic modulo 2. These circumferential waves
do not propagate in the z direction in the axis of the cylinder
or along its length. Guided circumferential waves have been
studied in Refs. 4 and 5 and were first introduced by
Viktorov6,7 as the natural extension of Lamb waves from a
flat plate to a curved plate. Liu and Qu8 consider the time
harmonic analysis of a perfect hollow cylinder, while Valle
et al.4 examine the time harmonic behavior of a doublelayered, hollow cylinder. None of these studies consider transient circumferential waves propagating in a cracked cylindrical component.
This paper examines the propagation of transient
guided circumferential waves in a metallic steel hollow

0001-4966/2001/110(3)/1282/9/$18.00

2001 Acoustical Society of America

FIG. 1. Specimen geometry, taking


advantage of symmetry of hollow cylinder, and simulating laser interrogation with a normal point load/point receiver system for backscattered energy
detection.

cylinder that contains a crack, using these guided waves to


both locate and size the crack. The crack is sized using a
scattering formula first developed by Auld,9 which is modified in the current work to analyze transient non-timeharmonic signals. The crack is then located by operating on
the backscattered signal with a time-frequency digital signal
processing DSP technique, and then comparing these results to those obtained from a perfect cylinderone without
a crack. All the time signals guided circumferential waves
presented in this work are generated using a commercial
FEM code ABAQUS/Explicitexcept for accuracy checks
done by comparing experimental and numerical results.
While much fundamental work has been done in the past to
develop numerically efficient methods to calculate the transient response of waveguides both perfect and flawed,10,11
the FEM was chosen here for its robustness, accuracy and
convenience.
It is important to note that no effort is made in this work
to selectively restrict the number of guided modes propagating in the cylinder.
One objective of this work is to demonstrate the effectiveness of using sophisticated DSP techniques to describe
guided, broadband, and multi-mode circumferential waves in
imperfect cylinders. Specifically, these DSP techniques are
used to examine the effect of scattering on dispersive waves,
showing it is possible to characterize cracks systematically
and accurately by quantifying this scattering effect.
In this text, the term wave designates any ultrasonic signal that propagates into the annular structure under investigation. The term mode designates an ultrasonic signal fully
represented in a time harmonic fashion by a single curve on
the dispersion curves of the structure where the mode propagates such as, the familiar Lamb modes a 0 and s 0 . Therefore, a mode is always a wave, but a wave can consist of
multiple modes since, as an example, the a 0 and s 0 modes
can co-exist for a given frequency range.
II. FEM MODEL OF THE SCATTERING OF GUIDED
CIRCUMFERENTIAL WAVES CAUSED BY A
CRACK

In the past, the computational cost of modeling highfrequency wave propagation problems was prohibitively
J. Acoust. Soc. Am., Vol. 110, No. 3, Pt. 1, Sep. 2001

high. It has been shown recently that the FEM is capable of


modeling such problems accurately and efficiently. This
computer power increase makes the FEM an excellent
method for solving wave propagation problems that are intractable by analytical methods.12,13
This numerical study only models a portion half of the
cylinder, because the signal processing techniques only make
use of the backscattered data and thus the whole circumference need not be modeled. The backscattered data will reach
the receiver before the transmitted data, and therefore the
backscattered data can be windowed out. Also, FE simulations show that most of the signal attenuates by the time it
has traveled 360. Therefore, there is no need to model the
complete cylinder. As such this geometry may be thought of
as an infinitely long curved plate. ABAQUS/Explicit is used
for all numerical calculations. Convergence is reached when
the solution obtained is mesh-invariant the solution remains
the same when the mesh size is decreased. As a further
check, solutions for the perfect cylinder are compared to the
normal modal expansion NME calculations of Liu and
Qu14 see Fig. 2, and to experimentally generated signals
using laser ultrasonics,5 shown in Fig. 3. The excellent
agreement between these solutions confirms the accuracy of
these FEM models. A typical half cylinder mesh the dimensions of the cylinder used in this research are shown in Table
I has 4000 elements for loads of 0.5 MHz center frequency. This number of elements is obtained by successive
refinement of the mesh until mesh invariance is obtained.
The element size is 15 times smaller than the wavelength of
the nondispersive shear wave at the center point of the
bandwidth 0.5 MHz and nearly 7 times smaller at the upper
end of the bandwidth 1.5 MHz. While these ratios are low
at the higher frequencies, mesh invariance and the numerical
validation with Fortran show that they are sufficient for
ABAQUS/Explicit.
Each element is a 4-node plain strain continuum element
CPE4R. Such an element provides a second-order interpolation, with reduced integration and hourglass control hourglassing is a numerical phenomena by which a zero-energy
mode propagates through and spoils the solutionsee the
ABAQUS Theory manual,15 Sec. 3.1.1, for more details.
Valle et al.: Crack characterization using guided waves

1283

TABLE I. Material and geometry data for the thin steel annulus.

FIG. 2. Comparison between wave forms from Abaqus/Explicit and


FORTRAN normalized units Ref. 12.

Each node has 2-degrees of freedom plane strain assumption.


Cracks in the cylinder are modeled by releasing nodes
along the cylinders wall and are assumed to have perfectly
smooth faces, subjected to stress free boundary conditions.
Such mesh discontinuities are infinitely thin, are a good representation of a fatigue crack as opposed to notches, and
allow for a variable crack length. In the far field which is the
domain of interest here. For near field calculations, singular
elements must be used to accurately capture the singularity at
the crack tip. It is important to note that the inverse problem
solved in this study does operate under the assumption that
the shape of the crack is known a` priori. In addition, while

cT

cL

0.2817

3120 m/s

5660 m/s

5.08 cm

5.38 cm

0.944 24

crack closure is of clear concern in practical cases, it will not


be addressed in this work.
A finer mesh is needed to accurately model cylinders
with smaller cracks i.e., of length smaller than 20% of the
wall of the waveguide. Convergence is usually obtained after increasing the number of elements in both the thickness
and the waveguide axis directions. Therefore, the computational models presented here can detect cracks of length no
smaller than 10% of the wall, that is, no smaller than 0.3 mm
for a load with 0.5 MHz center frequency. More powerful
computers can capture scattering due to even smaller cracks
because they can accurately resolve the bigger meshes
needed for loads of higher center frequencies.
The load in all cases is a point load, applied normally to
the cylinders surface. This load has a frequency range of 0.2
to 1.5 MHz, with most of the energy centered near 0.5 MHz.
The solution of such a FEM model takes a few minutes on a
Pentium III with a 433 MHz clock and 254 MB RAM.

III. AULDs FORMULA

Aulds formula9,16 is developed using a two transducer,


through transmission system applied to the structure to be
interrogated. Transducer 1 produces an incident field of
power P, and transducer 2 is the receiver see Figs. 4a and
b. The ratio of received electrical signal strength, E II , to
incident electrical signal strength, E I , is denoted by . The
change in the ratio, , due to a single scatterer a crack in
this case, is proportional to the reflection coefficient R and is
given by Aulds formula:

E II flaw E II noflaw / E I flaw.

This formula may be simplified for the case of backscattering, and with the hypothesis that the signals are time harmonic:

i
4P

2 1
1 2
k j uk k j uk n j

dS,

where S is an arbitrary surface, which surrounds the scatterer


the crack in this case, j and k are dummy indices the summation convention is used here and n j is the unit outward
normal of S. In Eq. 2 the terms with superscript 1 relate to
the fields in the absence of the scatterer, while terms with the
superscript 2 relate to the fields in the presence of the scatterer. For the case where the scatterer is a traction free crack,
Eq. 2 is simplified to


FIG. 3. Comparison between a laser-based experimental signal and the
equivalent ABAQUS/Explicit signal.
1284

J. Acoust. Soc. Am., Vol. 110, No. 3, Pt. 1, Sep. 2001

i
4P

k1j u k2 n l dS,

3a

where A is the crack area and


Valle et al.: Crack characterization using guided waves

FIG. 4. a Aulds formula general scattering geometry. b Aulds formula


backscattering case.
FIG. 5. Aulds formulascheme of FEM computations.

u k2 u k x
2 u k x 2 .

3b

The two previous equations are used to calculate the backscattering from a surface crack at the inner radius of the
cylinder.
Since the stress and displacement signals obtained from
ABAQUS are transient and not time harmonic, they cannot
be multiplied directly into the integrand of Eq. 3. The typical procedure17,18 involves applying the fast Fourier transform FFT to the transient displacement and stress signals
the signals calculated with ABAQUS first, then multiplying
them element by element, array multiplication and not vector multiplication, repeating the steps for each new value,
and finally adding everything. Once the sum is obtained, it
has to be multiplied by i element by element again and
the resulting output is Aulds formula in the frequency domain. In addition, the inverse FFT has to be applied if a time
domain result is desired. This procedure is lengthy and requires careful attention to the proper application of the FFT.
In particular, zero-padding the time-domain data is absolutely necessary. Otherwise, the final signal violates causality.
An interesting alternative to the frequency-domain formulation is to adapt Eq. 3 so that it can be applied directly
to time-domain signals. It is well known that the convolution
in the time domain is equivalent to the multiplication in the
frequency domain. In fact, many convolution programs use
that property directly in their codes: they first zero-pad the
data, apply the FFT, array multiply, and apply the inverse
FFT to get the final outcome.19 Equivalently, MATLAB uses
J. Acoust. Soc. Am., Vol. 110, No. 3, Pt. 1, Sep. 2001

a filter-based approach whereby the structures of the data are


cross multiplied:

y n f n *g n

f nm .g m .

The advantage of using the cross multiplication, as done


in MATLAB, is that much fewer operations are required
than in the alternative FFT-based formulation and that the
output is directly in the time domain. The only additional
step that is required for agreement with Eq. 3 is a time
differentiation, since multiplying by i in the frequency domain is equivalent to differentiating with respect to time in
the time domain.
The time-domain reflection coefficient, R, is therefore
calculated by using the following formula up to a multiplicative constant:
R

d
dt

n
u ni * i1
,

where n is the number of the element in consideration for


the stress tensor and of that elements bottom node on the
crack surface for the displacement field, * denotes convolution, and i1,2.
All the cracks considered in this study are radial e.g.,
vertical along the 90 angle, or the y axis. Therefore, the
stress tensor will always have a normal in the 0 angle direction, or the x axis see Fig. 5. Therefore, the j index is
always equal to 1.
Valle et al.: Crack characterization using guided waves

1285

wall. The slope also tends to flatten slightly as the crack


depth increases beyond 70%. Please note that mesh size the
number of elements needed to ensure convergence of the
ABAQUS model increases as the crack length decreases.
The quasilinear characteristic of the RMS plot shows
that Aulds method, combined with the FEM model, gives a
quantitative estimate of the crack length in a cracked steel
cylinder using guided circumferential waves. Moreover, this
method allows the sizing of cracks up to 300 m even
though the signal frequency is centered at 0.5 MHz and
never goes beyond 1.5 MHz. Note that a Rayleigh wave
with a wavelength of 300 m has a frequency of 9.7 MHz in
steel.
V. TIME-FREQUENCY REPRESENTATIONS TFRs OF
GUIDED WAVES
FIG. 6. Normalized RMS of FFT of Aulds reflection coefficient R versus
crack depth.

IV. REFLECTION COEFFICIENT VERSUS CRACK


LENGTH

Clearly, the longer the crack, the higher the overall level
of the resulting reflection coefficient. One way to obtain a
quantitative measure of this relationship is to plot the RMS
root-mean-square value of the frequency spectrum of the
reflection coefficient obtained via Aulds formula, for various
crack lengths. That is, once the FE model is used to calculate
the field around the crack, Aulds formula is then used to
predict the length of the crack with this FE data. The procedure involves a series of steps.
1 Construct a FEM model of the cylinder with a prescribed
crack length.
2 Solve for the pertinent displacement and stress fields
around the crack for the load case with a center frequency of 0.5 MHz.
3 Obtain the time-domain transient reflection coefficient,
R, from the procedure described in the preceding section
and by applying Eq. 5, up to an arbitrary multiplicative
constant.
4 Apply the FFT to R to get its frequency spectrum.
5 Get the RMS value of Rs frequency spectrum, and normalize it with respect to the RMS value of the incident
signal.
6 Plot the normalized RMS value with respect to crack
length, and repeat the entire procedure for other crack
lengths.
The resulting plot is shown in Fig. 6 for six different
crack lengths 10%, 20%, 30%, 50%, 70%, and 90% of cylinder wall. Figure 6 demonstrates that there is a quasilinear
correlation between the crack length and the overall value of
the reflection coefficient. For a perfect cylinder crack length
of 0% there should be no backscattered energy and thus the
normalized RMS should be 0. If the cylinder is completely
cut crack length of 100% the backscattered energy should
be equal to the incident energy. Figure 6 shows that the slope
is steeper for crack depths smaller than 10% of the cylinder
1286

J. Acoust. Soc. Am., Vol. 110, No. 3, Pt. 1, Sep. 2001

Now that a cracks length can be quantitatively measured, consider a procedure to determine the cracks location.
Once again the main challenge is the dispersive and multimode nature of guided waves. For example, consider a timedomain scheme that uses arrival times to determine the unknown location of a crack on a cylinders inner surface.
Since this scheme must be based on the arrival times of a
particular wave form feature, the dispersive nature of guided
waves presents a large source of error. The reason for this
potential breakdown is that the different frequencies of a
dispersive wave travel with different group velocities, changing the waves shape as it propagates. As a result, it is difficult if not impossible to track and identify the exact arrival
time of the same feature of a propagating guided wave.
In order to alleviate this problem, changes in the frequency content of the signal need to be tracked as a function
of time so that dispersion can be taken into account.
The time-frequency representation TFR of a signal is a
quantitative measure of how a signals frequency changes
with respect to time. A TFR is obtained by dividing a timedomain signal into a series of small pieces in time; each of
these pieces is windowed and then individually transformed
into the frequency domain. Popular transforms include wavelets, the short time Fourier transform STFT, and the
WignerVille distribution WV. Please recall from the preceding discussion that the identification of an individual
modes arrival time is very difficult if not impossible from
either the time-domain signal or its frequency spectrum, because the contributions from each mode in a multimode signal are not separable. However, the TFR enables the separation of the contribution of each mode, as a function of time
and frequency simultaneously.
One problem inherent to a TFR is the time-frequency
resolution limitation, that is, the impossibility to simultaneously have perfect resolution in both time and frequency.
One way to increase time-frequency resolution, is through
the reassignment or reallocation method.20 Reassignment is
not another TFR, but a way to reduce the spread of a TFR by
concentrating its energy to its the energys center of gravity. The reassignment method is not restricted to a specific
TFR, but can be applied to any time-frequency shift invariant
distribution of Cohens class.21 The energy density spectrum
of the reassigned STFT, called the reassigned spectrogram, is
Valle et al.: Crack characterization using guided waves

FIG. 7. a Comparison between the


reassigned spectrogram of a FEobtained guided circumferential wave
and the analytical dispersion curves, in
the time-frequency domain, of the
same cylinder both perfect. b
Methodology to locate the crack: comparison between the reassigned spectrogram for the backscattered energy
and the analytical dispersion curves in
the time-frequency domain.

selected for this study, since previous work21,22 shows that it


is extremely effective in capturing the dispersive nature of
guided, multimode ultrasonic signals in thin metallic plates.
Therefore, as was done in earlier studies for plates,22 it is
possible to compare analytical dispersion curves for the cylinder calculated in the time-frequency domain to a TFR
calculated from either experimentally or numerically FEM
generated guided circumferential wave signals. See Fig. 7a
for a comparison between the reassigned spectrogram of a
synthetic, i.e., numerical, signal from a perfect cylinder and
the analytical dispersion curves, calculated in the timefrequency domain, for that same perfect cylinder. The ladder effect described earlier in plates22 is clearly still present
for guided circumferential waves, and although the Rayleigh
mode is well represented by the TFR, the procedure breaks
down beyond 0.8 MHz because the FE models mesh is optimized for frequencies around 0.5 MHz.
J. Acoust. Soc. Am., Vol. 110, No. 3, Pt. 1, Sep. 2001

One of the greatest advantages of the proposed technique is that both numerical and experimental TFRs are not
limited to perfect geometries, while analytical dispersion
curves in general have this limitation. It is thus possible to
compare in the time-frequency domain analytically obtained
dispersion curves for a perfect cylinder to a reassigned
spectrogram obtained from a signal numerically calculated
or experimentally measured in a cracked cylinder, using
this comparison to quantify the cracks effect on these dispersion curves. This comparison offers a systematic procedure to locate a crack that does not require the selective
generation or detection of a particular mode. Applying the
reassigned spectrogram to backscattered ultrasonic energy
and comparing it to analytical dispersion curves for the cylinder calculated in the time-frequency domain provides an
excellent way to calculate the arrival times of specific modes
see Fig. 1 for the positioning of the source, the receiver and
Valle et al.: Crack characterization using guided waves

1287

the crack. A typical plot that results from the proposed procedure is given in Fig. 7b.
From Fig. 7b it is possible to determine:
which modes are affected by the crack,
for what times and what frequencies,
and even more importantly, by how much time is the arrival of each mode changed by the modes interaction with
the crack, for any given frequency.
This information is then used to determine the cracks
location. The next section will illustrate how this information
is extracted from the signals.
VI. CRACK LOCATION USING THE REASSIGNED
SPECTROGRAM

This portion of the study develops FE models for two


cases of the same steel cylinder, one with a crack length of
10% of the cylinders wall on the inner surface and a load
of center frequency of 0.5 MHz, and the second is the associated perfect cylinder no crack. The time-domain signal
backscattered from this crack is determined by subtracting
the signal predicted with the cracked cylinder model from
the signal of the perfect cylinder i.e., the model without a
crack, noting that each signal has the same source and receiver locations. The reassigned spectrogram is then applied
to this backscattered signal, and this TFR is compared to the
perfect cylinders analytical dispersion curves calculated,
in the time versus frequency domain, for the same source
and receiver locations. This comparison illustrates which
modes are affected by the crack, as well as enabling the
calculation of the time delay between the backscattered and
the corresponding analytical dispersion curves see Fig.
7b. The reason for this time delay is that the backscattered
signal only starts after the incident signal has propagated
from the source to the receiver as in the perfect cylinder
case and then to the crack and then back to the receiver.
Therefore, the backscattered signal is zero for all times prior
to the time that it takes to return to the receiver.
This time delay is given by the ratio between the group
velocity of each specific mode, for a specific frequency, and
twice the distance between the receiver and the crack. Note
that there is a distortion in the backscattered reassigned spectrogram, when compared to the analytical dispersion curves,
because the group velocity of most modes changes with
frequencythis causes a change in the time delay, with frequency. As a result, the shape of the backscattered spectrogram looks slightly different from the shape of the corresponding analytical dispersion curves for the perfect cylinder
see Fig. 7a.
This distortion makes it difficult to identify exactly
which feature of a given backscattered mode, for a specific
frequency, corresponds to what feature of the analytical dispersion curve for that mode and for what frequency. For
example, in Fig. 7a, the highest peak in the second mode
occurs at 0.82 MHz in the analytical dispersion curve, but
finding this exact same feature in the backscattered spectrogram for the second mode is impossible since the peak is
now spread over a small frequency range 1 to 1.2 MHz.
Therefore, since the group velocity depends on frequency,
1288

J. Acoust. Soc. Am., Vol. 110, No. 3, Pt. 1, Sep. 2001

FIG. 8. Comparison between predicted distances receiver to crack lower


and upper bound and the true distance, for the ring cracked at 10%.

the time delay between the reassigned spectrogram of a


backscattered mode and its dispersion curve will also be hard
to determine accurately. The only mode for which the preceding statement is not true is the Rayleigh mode, because its
group velocity is constant for almost all frequencies, so the
frequency distortion between the Rayleigh modes analytical
dispersion curve and its reassigned spectrogram is nearly
nonexistent beyond 0.4 MHz, shown in Fig. 7a. The time
delay between the backscattered Rayleigh mode from the reassigned spectrogram and its predicted analytical arrival,
times the group velocity of the Rayleigh mode, gives twice
the distance between the receiver and the crack.
Since the scattered Rayleigh mode, as calculated with
the reassigned spectrogram, is not an infinitely thin line as is
clear from Figs. 7a and b, two distances are always calculated, one corresponding to the lowest bound in time first
time, and one corresponding to the upper bound in time
second time. At 0.5 MHz, the Rayleigh wave can be assumed to be almost nondispersive, thus if it reaches the receiver in a faster time, it has traveled a comparatively shorter
distance. Hence the lower bound on the distance to, or
location of, the flaw. Conversely the second and longer time
corresponds to a longer distance traveled by the backscattered signal to the receiver. Hence the lower bound on the
distance to the flaw. These distances can then be plotted
against the true distance, i.e., the actual location of the crack.
The comparison between those three distances is given in
Fig. 8. There is very good agreement between them, even for
this small 10% of wall thickness crack.
The proposed methodology also works well in the
nearfield of the crack or source, although it is more difficult
to accurately determine separation times between backscattered and analytical modes in this case; for example, the
analytical modes will tend to be compressed all together toward the zero time axis if the receiver is close to the source.
Since the method is based on a graphical calculation of time
Valle et al.: Crack characterization using guided waves

delay, it is essential to have a definite separation between the


analytical dispersion curves and their backscattered counterparts.
More importantly, the Rayleigh mode is not always the
mode most sensitive to the crack. Indeed, in Fig. 7b, one
can see that for certain frequency ranges for example, from
1 MHz on, the backscattered contribution from the Rayleigh
mode simply vanishes. This means that for frequencies
above about 1 MHz, the Rayleigh mode does not see the
crack this makes sense since the Rayleigh mode travels on
the outside radius while the crack originates from the inside
surfacetherefore, as the frequency increases the Rayleigh
mode is increasingly less likely to see the crack. Only the
higher order modes in this case the second mode see the
crack at that particular frequency as was first shown in Valle
et al.4this is clearly visible in Fig. 7b. Unfortunately,
frequency distortion is a problem for these higher modes.
This frequency distortion makes it difficult to use higher order modes to determine the location of the crack, unless one
can select a frequency range with limited dispersion, so that
the group velocity can be calculated accurately. Typically
this frequency range exists, but it may be prohibitively high
from a computational cost perspective. Please note that dispersion curves and FE signals tend to be costly for frequencies beyond 3 to 4 MHz.
This problem the Rayleigh mode not being sensitive to
cracks only occurs for very small cracks. The reassigned
spectrogram applied to time-domain signals created by a 0.5
MHz signal is clearly sufficient to locate this 300 m crack
10% of the cylinders thickness.
VII. SUMMARY AND CONCLUSIONS

This paper uses the FEM to model the propagation of


guided circumferential waves in a cracked cylinder and presents sophisticated digital signal processing algorithms that
characterize the waves interaction with this crack. Specifically, this paper presents a technique to size the crack, and
another to locate it.
For sizing the crack, a time domain analysis based on
Aulds scattering formula for backscattered signals is used.
By using the Aulds formula to calculate the RMS value of
the frequency spectrum of the reflection coefficient for various crack lengths, it is shown that there is a quasilinear correlation between the crack depth and the overall RMS value
of the reflection coefficient R. The RMS value of R increases
with crack depth, and provides a quantitative means of characterizing cracks down to about 10% of the thickness of the
cylinder. As a result, Aulds method combined with the FEM
model gives a quantitative estimate of the crack depth from a
measured signal.
To locate the crack, the reassigned spectrogram is presented as a digital signal processing algorithm that accurately
captures a TFR of guided circumferential waves over a large
frequency range. The reassignment method solves the timefrequency resolution problem inherent to using Fourier-based
TFRs such as the spectrogram by redistributing the energy
content of each mode to its center of gravity, with respect to
time and frequency, and therefore clearly separates the contribution of each mode within the scattered signal. Applying
J. Acoust. Soc. Am., Vol. 110, No. 3, Pt. 1, Sep. 2001

the reassigned spectrogram to backscattered data, and comparing this to the analytical dispersion curves for the perfect
cylinder provides an accurate measure of the time delay, per
mode, between the incident signal of a perfect waveguide
and its backscattered counterpart if a crack is present.
Through this time delay and the group velocity of the mode
of interest at a specific frequency, the distance between the
receiver and the crack can be calculated. This method can
characterize cracks and notches with lengths as small as 10%
of the waveguide wall. Computational power needs to be
increased if even smaller sizes are of interest.
It is important to note that the results presented in this
paper, both for crack sizing and crack locating, are dependent
on the frequency of the input signal. Both methods Aulds
formula and the reassigned spectrogram are applied to signals created with an input with a frequency range between
0.2 and 1.5 MHz with most of the energy centered around
0.5 MHz but they can detect cracks down to 300 meven
though the wavelengths of these signals is much greater than
300 m. Therefore, the need for high frequency signals to
detect small cracks is significantly reduced by using these
techniques.
Also, multiple cracks can be detected using those techniques. The reassigned spectrogram will locate them by
showing a series of time delays. Aulds formula as presented
here with no restriction on the duration of the signals can
only show a combined total of damage accumulation
however, if it is used in combination with the reassigned
spectrogram, the signals used in Aulds formula can be restricted to the pertinent time of flight and therefore can then
characterize each crack singly, in the same fashion presented
earlier in this paper.
ACKNOWLEDGMENTS

This work is supported by the Office of Naval Research


M-URI Program Integrated Diagnostics Contract No.
N00014-95-1-0539. Two Amelia Earhart Fellowships from
the Zonta Foundation, and support from the Deutscher Akademischer Austausch Dienst DAAD are also gratefully acknowledged by the first and second authors, respectively.
1

P. B. Nagy, M. Blodgett, and M. Godis, Weep hole inspection by circumferential creeping waves, NDT & E Int. 27, 131142 1994.
2
D. E. Chimenti, Guided waves in plates and their use in material characterization, Appl. Mech. Rev. 50, 247284 1997.
3
D. N. Alleyne and P. Cawley, The long range detection of corrosion in
pipes using lamb waves, Rev. Prog. Quant. Nondestr. Eval. 14, 2073
2080 1995.
4
C. Valle, J. Qu, and L. J. Jacobs, Guided circumferential waves in layered cylinders, Int. J. Eng. Sci. 37, 13691387 1999.
5
C. Valle, Guided circumferential waves in annular structures, Ph.D.
thesis, Georgia Institute of Technology, Atlanta, Georgia, 1999.
6
I. A. Viktorov, Rayleigh and Lamb Waves Plenum, New York, 1967.
7
I. A. Viktorov, Rayleigh-type waves on a cylindrical surface, J. Acoust.
Soc. Am. 4, 131136 1958.
8
G. Liu and J. Qu, Guided circumferential waves in a circular annulus, J.
Appl. Mech. 65, 424 430 1997.
9
B. A. Auld, Acoustic Fields and Waves in Solids, 2nd ed. R. E. Krieger,
Melbourne, FL, 1990, Vols. 1 and 2.
10
J. Zhu, A. H. Shah, and S. K. Datta, Transient response of a composite
plate with delamination, J. Appl. Mech. 65, 664 670 1998.
11
S. K. Datta, T. H. Ju, and A. H. Shah, Scattering of an impact wave by a
crack in a composite plate, J. Appl. Mech. 59, 596 603 1992.
Valle et al.: Crack characterization using guided waves

1289

12

F. Moser, L. J. Jacobs, and J. Qu, Modeling elastic wave propagation in


wave guides with the finite element method, NDT & E Int. 32, 225234
1999.
13
M. J. S. Lowe, R. E. Challis, and C. W. Chan, The transmission of Lamb
waves across adhesively bonded lap joints, J. Acoust. Soc. Am. 107,
13331345 2000.
14
G. Liu and J. Qu, Transient wave propagation in a circular annulus
subjected to transient excitation on its outer surface, J. Acoust. Soc. Am.
104, 12101220 1998.
15
ABAQUS-Standard version 5.6 Users Manuals, Theory Manual, Example Problems Manual, and ABAQUS-Post Manual, 1996.
16
B. A. Auld, General electromechanical reciprocity relations applied to
the calculation of elastic wave scattering coefficients, Wave Motion 1,
310 1979.
17
M. J. S. Lowe, P. Cawley, J-Y. Kao, and O. Diligent, Prediction and
measurement of the reflection of the fundamental anti-symmetric Lamb

1290

J. Acoust. Soc. Am., Vol. 110, No. 3, Pt. 1, Sep. 2001

wave from cracks and notches, Rev. Prog. Quant. Nondestructive Eval.
19A, 193200 2000.
18
M. Lowe, Characteristics of the reflection of lamb waves from defects in
plates and pipes, Rev. Prog. Quant. Nondestr. Eval. 17, 113120 1997.
19
W. H. Press, B. P. Flannery, S. A. Tenkolsky, and W. T. Vetterling, Numerical Recipes Cambridge University Press, Cambridge, 1988.
20
F. Auger and P. Flandrin, Improving the readability of time-frequency
and timescale representations by the reassignment method, IEEE Trans.
Signal Process. 43, 1068 1098 1995.
21
M. Niethammer, Application of time frequency representation to characterize ultrasonic signals, M.S. thesis, Georgia Institute of Technology,
Atlanta, Georgia, 1999.
22
M. Niethammer, L. J. Jacobs, J. Qu, and J. Jarzynski, Time-frequency
representation of lamb waves using the reassigned spectrogram, J.
Acoust. Soc. Am. 107, L19L24 2000.

Valle et al.: Crack characterization using guided waves

Вам также может понравиться